Sie sind auf Seite 1von 5

JOURNAL OF PROPULSION AND POWER

Vol. 23, No. 5, SeptemberOctober 2007

Technical Notes

Downloaded by SATISH DHAWAN SPACE CENTRE (ISRO SHAR CTR) on March 13, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.23748

TECHNICAL NOTES are short manuscripts describing new developments or important results of a preliminary nature. These Notes should not exceed 2500
words (where a gure or table counts as 200 words). Following informal review by the Editors, they may be published within a few months of the date of receipt.
Style requirements are the same as for regular contributions (see inside back cover).

Burning-Rate Calculations
of Wide-Distribution Ammonium
Perchlorate Composite Propellants

Muhammad Mazhar Iqbal and Wang Liang


Northwestern Polytechnical University,
710072 Xian, Peoples Republic of China

Subscripts

b
j

=
=

ox
PF
p
0

=
=
=
=

associated with the binder


number of the oxidizer particle-size fraction in the
propellant
associated with the oxidizer
associated with the primary ame
associated with the propellant
represents the total or initial conditions

DOI: 10.2514/1.23748

I. Introduction

MMONIUM perchlorate (AP) used in composite propellants


consists of various blends of particle sizes for high loading
density and stable combustion. Three or four particle sizes are
commonly used in propellants, which are sufcient to come close to a
viscosity optimum for a specic solid loading. Most of the practical
propellant formulations have a wide-distribution multimodal
oxidizer blend containing both very large (200 to 400 m) and
very small (1 to 25 m) AP particles. The oxidizer particle size and
particle-size distribution of various blends of oxidizer have a
considerable effect on the burning rate of the propellant. The general
need is to describe the burning rate of these propellants in terms of
oxidizer particle-size distribution. The objective of this research is to
describe a combustion model for the determination of burning rates
and combustion properties for multimodal AP composite
propellants.
Combustion models based on the multiple-ames concept, such as
the BecksteadDerrPrice (BDP) model [1], petite ensemble model
(PEM) [2], Cohen and Strand model [3], and Beckstead models [4
6], have been able to predict the ballistic properties (burning rate and
pressure exponent) of various polydisperse propellants to within
about 10%. These models seek to simplify the statistically threedimensional propellant geometry by unimodal oxidizer particles of a
representative size surrounded by some binder layer of an equivalent
extent. The inuence of the size distribution on the average burning
rate is predicted by integrating the results of independent analysis of
each particle size. These models, however, are not able to predict the
burning-rate behavior of the wide-distribution propellants.
Wide-distribution propellants have unique combustion mechanisms at the propellant surface. Studies have shown that a thin layer
of liquid binder covers portions of the exposed AP surface, which is a
dominant combustion mechanism for wide-distribution solid
propellants [7]. This binder covering reduces the oxidizer burning
rate, and the difference in the measured and the predicted burning
rates is related to the extent of oxidizer surface covered with binder.
The oxidizer-to-fuel ratio in these propellants plays a dominant role
in combustion. These considerations were accounted for while
formulating the model for burning-rate calculations.
The basic modeling approach is based upon the BDP multipleames concept. The following changes were incorporated in the
original model to achieve the objective:
1) The PriceBoggsDerr (PBD) model [8,9] for AP deagration
is adopted with modied input parameters, which provides a realistic
means for examining and understanding the self-deagration
properties associated with solid propellants. The model includes
variable specic heat capacity and conductivity and describes selfdeagration as a function of pressure and initial temperature. As a
result, the temperature and pressure dependence of burning rates of

Nomenclature
Afh
As
Bfh
b
cg
cs
Dox
Es
k
m
O=F
P
Q
Qfuel
QL
R
r
S
Tox
Ts
T0
XD
X

F

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

ox
p

=
=

Hs

g



=
=
=
=
=

average ame-height factor with respect to the oxidizer


kinetic prefactor for surface decomposition
average ame-height factor for the binder
characteristic surface dimension
specic heat capacity of the oxidizer combustion gases
specic heat capacity of the solid
oxidizer particle size
activation energy for surface decomposition
rate constant
mass ow rate
oxidizer/binder ratio
pressure
heat released in the ame
heat of the binder decomposition
net heat released at the oxidizer surface
molar gas constant
burning rate
surface area
oxidizer monopropellant ame temperature
surface temperature
initial propellant temperature
diffusion-ame height
ame standoff distance
mass fraction in the propellant
fraction of the oxidizer that reacts in the diffusion
ame to heat the oxidizer
fraction of the oxidizer that is used to heat the oxidizer
fraction of the oxidizer not involved in the condensed
phase reaction
latent heat of the solid oxidizer
ame reaction order
gas thermal conductivity
dimensionless ame height
density

Presented as Paper 1161 at the 44th AIAA Aerospace Sciences Meeting


and Exhibit, Reno NV, 912 January 2006; received 9 March 2006; revision
received 13 March 2007; accepted for publication 30 March 2007. Copyright
2007 by the American Institute of Aeronautics and Astronautics, Inc. All
rights reserved. Copies of this paper may be made for personal or internal use,
on condition that the copier pay the $10.00 per-copy fee to the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include
the code 0748-4658/07 $10.00 in correspondence with the CCC.

Research Scholar, College of Astronautics.

Professor, College of Astronautics.


1136

J. PROPULSION, VOL. 23, NO. 5:

composite propellants can better be predicted when compared with


the previous models.
2) The oxidizer and binder are assumed with separate surface
temperatures in the current model. Separate energy balances are
taken for the oxidizer and binder to determine their respective surface
temperatures.
3) The model is extended to incorporate the multimodal oxidizer
sizes using a pseudopropellant approach following the Cohen and
Strand model [3]. The burning rate of each pseudopropellant is
calculated independently, using the monomodal model. The
aggregate propellant burn rate is determined from all the mass ow
rate and surface-area contributions of all the pseudopropellants.

8.5

r , mm/s

8
7.5
7
6.5
6
5.5
5
0

Fig. 1

Experimental Work

Ammonium perchlorate (oxidizer), hydroxyl-terminated polybutadiene (HTPB) (binder), aluminum (fuel), toluene diisocyanate
(curing agent), and some common propellant additives were selected
as the propellant ingredients for formulation experimentations. Three
to four particle sizes of AP were used in varying proportions to study
the effect on burning rate and pressure exponent. A sufcient
quantity of all the raw materials was prepared and homogenized to
ensure a uniform batch. A small-scale vertical mixer was used for
mixing the propellant slurry and later the scale was enhanced to a
larger mixer to observe the effect. Several mixes of propellant slurry
were prepared to observe the effect of the formulation variables.
Propellant slurry was cast in rectangular cartons under vacuum and
cured at 70 C in an oven for three days. Samples were prepared from
the cured propellant and were tested for burning rate, pressure
exponent, temperature sensitivity of burning rate, and various
mechanical properties (stress, strain, and modulus). The results
obtained (statistical mean value of ve test samples) from these tests
are summarized and discussed next.
The effects of varying the mixing ratio of coarse and ne AP and
the catalyst on propellant burning rates were studied. In the rst step,
trimodal AP compositions were varied but the total contents of AP
remained unchanged. Three types of spherical and one type of
nonspherical AP were used, referred to as types 1, 2, 3, and ne,
respectively. The mass average diameters d43 of types 1, 2, and 3
were measured as 340, 250, and 135 m, respectively. The average
diameter d43 of the nonspherical (ne) AP was measured as 10 m.
The proportions of various AP fractions with the measured burning
rate and pressure exponent are given in Table 1.
In formulations AH-101 and 104, types 1, 3, and ne AP were
used. Formulation AH-105 was reproduced with the same
compositions as that of no. 104, but replacing type 1 AP with
type 2. This change resulted in increasing the burning rate and
reducing the pressure exponent of the propellant, which is according
to the expected trend. The rest of the propellant formulations were
prepared by using type 2 AP as the coarse fraction. In formulations
AH 106108, the ne-AP proportion was increased from 525%,
which was further increased up to 35% in formulation leading to
no. AH-113. Formulation AH-110 was just the repetition of no. 108
to verify the results. A good reproducibility of the results was
observed from nos. 108 and 110.
The results of the burning rate versus ne-AP contents are plotted
and shown in Fig. 1. It can be seen from the gure that the burning

10

20

30

P = 5.5 Mpa
8

P = 6.0 Mpa

7
6
5
0

50

100

150

200

AH-101
AH-104
AH-105
AH-106
AH-107
AH-108
AH-110
AH-112
AH-113

250

300

Average AP diameter d 4 3, m
Fig. 2 Change in the burning rate with the AP mean diameter.

rate rises with the increase in ne-AP contents. The rise becomes
quicker when ne-AP contents are increased from approximately
20% onward. This effect is more clearly visible from Fig. 2, in which
the mass average diameter d43 of the three AP fractions is plotted
against the burning rate at 5.5 and 6.0 MPa.
The burning rate decreases noticeably with d increasing, but at
large d, the dependence of burning rate on d becomes weak. Such
behavior of burning-rate dependence on d was also observed in
previous studies with bimodal oxidizer at stoichiometric oxidizer/
fuel ratios [10,11].

III. Combustion Mechanisms of Wide-Distribution


AP Propellants
Wide distribution denotes a multimodal oxidizer blend containing
both very large and very small oxidizer particles. The size difference
allows higher levels of oxidizer-to-fuel ratio for the propellant to be
achieved. The general need is to describe the burning rate of these
propellants in terms of oxidizer particle-size distribution and binder
composition. Combustion models based on the multiple-ames
concept, such as BDP and the PEM, were able to predict the ballistic
properties of previous polydisperse propellants to within about 10%.

Table 1 Various proportions of coarse and ne AP with burning rate and pressure exponent
Formulation code

40

Fine AP, %
Variation in propellant burning rate with ne AP.

r , mm/s

Downloaded by SATISH DHAWAN SPACE CENTRE (ISRO SHAR CTR) on March 13, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.23748

II.

1137

TECHNICAL NOTES

AP, %
Fine

Type 3

Type 2

Type 1

3.5
11.5
11.5
5.0
15.0
25.0
25.0
30.0
35.0

24.0
28.0
28.0
31.5
26.5
21.5
21.5
19.0
16.5

28.0
31.0
26.0
21.0
21.0
18.5
16.0

40.0
28.0

r (6.0 MPa), mm/s

Pressure exponent n

5.81
6.02
6.44
6.49
6.39
6.93
6.88
7.48
8.74

0.420
0.398
0.387

0.400
0.430
0.430

0.530

Downloaded by SATISH DHAWAN SPACE CENTRE (ISRO SHAR CTR) on March 13, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.23748

1138

J. PROPULSION, VOL. 23, NO. 5:

The inuence of size distribution on the average rate is predicted by


integrating the results of independent analysis of each particle size.
These models, however, are not able to predict the burning-rate
pressure dependence of wide-distribution propellants.
Comparison between measured and predicted burning rates from
the previous studies are given in Fig. 3. Incrementally increasing the
replacement of ne particles with the coarse particles in the trimodal
AP propellant resulted in a decrease in measured burning rate.
Results of PEM calculations are close to the measured burning rates
at lower concentrations of coarse particles, but miss the decrease in
rate as the coarse fraction is increased. Two possible phenomena that
could cause this difference are 1) control of the ballistics by the neAP/HTPB (pocket propellant) matrix and 2) local intermittent
burning related to the mass fraction of the coarse oxidizer particles.
Because of the fuel-rich nature of the ne particle combustion, these
propellants also exhibit extreme sensitivity to changes in binder
composition.
Wide-distribution propellants have unique combustion mechanisms. The ballistic results of these propellants reveal anomalous

Tsox  T 

TECHNICAL NOTES

Similarly, the binder mass ow rate is given by


mb  b rb  Asb exp

B.

Separate Energy Balance for Oxidizer and Binder

mox Sox cs Tsox  T   Hs  QL


 ox F mox Sox  mb Sb QPF expPFox 
 ox 1  F mox Sox Qox expox 

Mass Flow Rates

The mass ow rates of the oxidizer and binder depend solely on


their respective surface temperatures through the Arrhenius
pyrolysis law. The mass ow rate of the oxidizer for a given surface
temperature is calculated by

mox  ox rox  Asox exp

Esox
RTsox


(1)

(3)

The surface temperature of the oxidizer is calculated from the


preceding equation and is given by

(4)

Similarly, the energy balance for the binder is taken as


mb Sb cs Tsb  T   Qfuel
 1  ox mox Sox  mb Sb QPF expPFb 

(5)

Equation (5) is used to calculate the ox which is given by


ox  1 

C.

mb Sb cs Tsb  T   Qfuel


mox Sox  mb Sb QPF expPFb 

(6)

Binder Regression Rate

The binder regression rate is calculated by continuity of the


propellant:
 

(7)
mb Sb  mox Sox b
ox

IV. Combustion Modeling


The combustion model is based on the BDP multiple-ames
concept [1]. Separate surface temperatures for the oxidizer and
binder were assumed following Beckstead and McCarty [5]. A
complete description of the model is given in [13]. The key equations
used in the model are summarized next.

(2)

The energy balance for the AP is written as

ox F mox Sox  mb Sb QPF expPFox   ox 1  F mox Sox Qox expox  Hs  QL

cs mox Sox
cs

burning rates. High-speed photographs and examination of


extinguished propellant surfaces have shown evidence of molten
liquid binder owing on the propellant surface. This was indicated by
the beads of binder owing off the burning surface in the windowbomb movies and thin layers of binder covering portions of the
oxidizer particles on the extinguished surface [7]. The binder ow
increases signicantly as the oxidizer-to-fuel ratio of the pocket
propellant is lowered.
The binder covering reduces the oxidizer burning rate and is the
main cause of differences between the predictions made with the
PEM combustion model and the experimental results for widedistribution propellants. The fraction of the AP burning surface
covered with the molten binder determines the magnitude of the
propellant burning-rate suppression. Experimental evidence
suggests that lowering the oxidize-to-fuel ratio of the pocket
propellant increases the binder-covering fraction. The bindercovering fraction also varies with the combustion pressure [12].

A.



Esb
RTsb

rb 

mb
b

(8)

This regression rate is used in the surface-area calculations to


determine Sox . Unlike the BDP model, it is done in iteration, because
rb is function of Sox in this calculation. The regression rate is then
used to determine Tsb from Eq. (5).
D.

Flame Heights

The BDP model used an approximation of the Williams and


BurkeSchumann diffusion ame to determine an effective
diffusion-ame height over the AP. The same basic approach is
used for the calculations of the diffusion-ame height. However, a
separate averaging method is applied to the diffusion-ame height
over the oxidizer and binder.

J. PROPULSION, VOL. 23, NO. 5:

1139

TECHNICAL NOTES

100

model (90/200 m)
experimental (90/200 m)
model(9/90 m)
experimental (9/90 m)

10
measurements

r, cm /s

r (6.8 MPa), mm/s

10

PEM calculations

1
35

40

45

50

55

Wt % 400- m AP

Fig. 3 Comparison of predicted and experimental measured burning


rates.

0.1

Downloaded by SATISH DHAWAN SPACE CENTRE (ISRO SHAR CTR) on March 13, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.23748

The dimensionless ame height over the oxidizer is given by



cs mp Xox
 XDox 
PFox 
g

XDox  Afh
XD

(9)

(10)


cs mp XPF
 XDb 
g

XDb  Bfh
XD

(11)

different rates. The model gives good results for nitramine


propellants associated with a signicant ignition delay time. The
Cohen and Strand [3] model is used for the purpose of present
calculations, which gave good results with modications described
earlier.
In the Cohen and Strand [3] approach, each modal component is
viewed as a pseudopropellant. Each pseudopropellant includes an
assigned portion of the binder. The burning rate is determined from
all of the mass ow rate and surface-area contributions by serially
building up pseudopropellant contributions:

(12)
mp;n

n
X

It turns out that the factor Bfh is much smaller than Afh , the average
ame-height factor with respect to the AP. Thus the binder is heated
by a relatively close-in ame, which is consistent with Burke
Schumann model solutions for ames that extend over the fuel.

is calculated by the following relationship:
XPF

XPF


mp
kPF PPF

(13)

Calculations of the diffusion-ame height XD and the heat-release


term used in Eqs. (4) and (6) are adopted from the BDP model and
elaborated in [14]. The characteristic surface dimension b that is used
to determine the diffusion-ame height XD is different from the BDP
model and is given later in this paper.
E.

100

Fig. 4 Results of bimodal AP/HTPB propellants compared with the


experimental data.

The dimensionless ame height over the binder is given by


PFb 

10
P, MPa

Soj  mox Sox jn  mb Sb jn  mp;n1

n1
X

Sj

(14)

r

mp;n
p

(15)

The rst two terms on the right side of Eq. (14) are the mass ow
contributions of the oxidizer and the binder for last pseudopropellant.
In the calculation procedure, j  n represents the coarsest-particlesize pseudopropellant. The characteristic surface dimension bj for
multimodal propellant calculations is given by [5]


Doxj
 =
(16)
bj  p 1  ox b
O=Fj
6

Multimodal Propellant Calculations

Several different approaches were used for building up the


component modal contributions to express the aggregate propellant
burning rate. The two most common approaches presented by Cohen
and Strand [3] and Beckstead [6,10] were explored during this study.
The original Cohen and Strand [3] modeling approach failed to
calculate the burning rates of AP propellants having a wide
distribution of particle sizes. In his later models, Beckstead [6,15]
used a time-averaging approach, assuming that the propellant burns
through alternate layers of binder and oxidizer at signicantly

The parameter O=Fj depends upon the apportioning of the


binder to the oxidizer among the pseudopropellants. Cohen and
Strand used a constant O=Fj value in their model. This was
changed and the effective O=F ratio for each pseudopropellant is
assigned according to their relative proportion in the propellant. This
approach affords the opportunity to predict more realistic burning
rates and accounts for wide-distribution phenomena. The surfacearea expressions used in the BDP mode are extended to incorporate
the multimodal propellants.

Table 2 Experimental and calculated results for trimodal aluminized propellants


Formulation

AP proportions 10/135/250/340 m

AH-101
AH-104
AH-105
AH-106
AH-107
AH-108
AH-112
AH-113
a

The pressure is 6.0 Mpa.

3.5/24//40
11.5/28//28
11.5/28/28/
5/31.5/31/
15/26.5/26/
25/21.5/21/
30/19/18.5/
3.5/16.5/16/
b

The pressure range is 410 MPa.

Burning rate,a mm=s

Pressure exponentb

Experimental

Calculated

Experimental

Calculated

5.81
6.02
6.44
6.49
6.39
6.90
7.48
8.74

5.10
5.58
5.62
5.28
5.82
6.38
6.67
6.96

0.420
0.398
0.387

0.400
0.430
0.430
0.530

0.361
0.358
0.387

0.412
0.485
0.520
0.562

1140

J. PROPULSION, VOL. 23, NO. 5:

TECHNICAL NOTES

burning rate on initial temperature is not computed in the present


study and work on that is still in progress. Because the model predicts
the combustion properties of practical propellant formulations to a
reasonable accuracy, overall propellant optimization can be
performed and, as such, signicant gains can be reached by
reducing the time and cost for new propellant formulation
development.

References

Downloaded by SATISH DHAWAN SPACE CENTRE (ISRO SHAR CTR) on March 13, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.23748

Fig. 5 Pressure dependence of the burning rate for trimodal AP/HTPB


formulations.

V. Results and Discussions


Calculations are performed for a bimodal propellant and compared
with the experimental data of King [16] and Miller [17,18] for the
verication of the model. These propellants are good choices
because they encompass a range of size distributions and AP
concentrations. Calculation results are compared with experimental
data of Kings [16] formulation having 82% AP (90=200 m, 1=5
ratio) and Millers [17,18] formulation having 80% AP (9=90 m,
1=1 ratio) and are shown in Fig. 4. The results show very good
agreement with the data in both cases. A rise in the slope of 9=90 m
propellant curve is attributed to the shift from the diffusion-amecontrol regime to AP ame control at about 10 MPa.
The model is extended to calculate the burning rates of widedistribution particle sizes and aluminized propellants. Our own
experimental data, which contain various formulation results of
aluminized AP/HTPB propellants [19], are used for the calculations.
These data are a good choice due to their versatile nature, having
trimodal AP with variable particle sizes and proportions, and they
provide the pressure exponent for most formulations. These data
represent the real practical formulations. The results of burning rate
and pressure exponent are compared and shown in Table 2. The
calculated results are in very good agreement with the relevant
experimental data. All of the results are within 20% deviation: 75%
results lie within 15% deviation, and 50% calculations are within
10% deviation. The statistics shows the excellent predictions for such
a practical wide-distribution aluminized propellant formulation. The
pressure versus burning-rate curve for all the formulations of [19] are
also plotted and shown in Fig. 5. This shows the good computational
capability of the model with respect to both pressure and particle-size
effects. Thus the desired purpose has been achieved.
All of the propellants were calculated at the initial temperature of
298 K. The effect of initial propellant temperature requires further
work, which is still in progress.

VI.

Conclusions

A model is presented for the combustion calculations of AP-based


composite propellants within the framework of the BDP model. The
model differs from the BDP model in the sense that it accounts for the
latest PBD model for AP deagration with modied parameters. A
separate energy balance is assumed for the binder to calculate
different surface temperatures of the oxidizer and binder. A
pseudopropellant approach is used for extension of the model to
multimodal AP propellants. New parametric values used in the PBD
model enable more widened and more realistic surface-temperature
calculations of oxidizer and better predictions of monopropellant
rates. Burning-rate calculations for various monomodal AP/HTPB
propellants were performed and compared with the experimental
data. Results are in good agreement. Then the model is extended to
multimodal propellants and results are compared for bimodal AP
propellant formulations that again show good agreement with the
data. Finally, the results are computed for the most practical
propellant formulations containing aluminum and trimodal AP in a
wide distribution of particle sizes. The results of burning rate and
pressure exponent are in very good agreement with the experimental
data. Results show the good capability of the model for determining
the effect of both pressure and AP particle size. The dependence of

[1] Beckstead, M. W., Derr, R. L., and Price, C. F., A Model of Composite
Solid Propellant Combustion Based on Multiple Flames, AIAA
Journal, Vol. 8, Dec. 1970.
[2] Glick, R. L., Distribution Functions for Statistical Analysis of
Monodisperse Composite Solid Propellant Combustion, AIAA
Journal, Vol. 14, Nov. 1976.
[3] Cohen, N. S., and Strand, L. D., An Improved Model for the
Combustion of AP Composite Propellants, AIAA Journal, Vol. 20,
No. 12, Dec. 1982, pp. 17391746.
[4] Beckstead, M. W., Derr, R. L., and Price, C. F., The Combustion of
Solid Monopropellants and Composite Propellants, Thirteenth
Symposium (International) on Combustion, Combustion Inst.,
Pittsburgh, PA, 1971, pp. 10471056.
[5] Beckstead, M. W., and McCarty, K. P., Modeling Calculations for
HMX Composite Propellants, AIAA Journal, Vol. 20, No. 1,
Jan. 1982, pp. 106115.
[6] Beckstead, M. W., A Model for Solid Propellant Combustion,
Eighteenth Symposium (International) on Combustion, Combustion
Inst., Pittsburgh, PA, 1981, pp. 175185.
[7] Frederick, R., Jr., and Osborn, J. R., Ballistic Studies of Wide
Distribution Propellants, 36th AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit, Huntsville, Al., AIAA Paper 20003318, July 2000.
[8] Price, C. F., Boggs, T. L., and Derr, R. L., The Steady State
Combustion Behavior of Ammonium Perchlorate and HMX, 17th
Aerospace Sciences Meeting, AIAA Paper 79-0164, Jan. 1979.
[9] Price, C. F., Boggs, T. L., and Derr, R. L., The Modeling of Solid
Monopropellant Deagration, 16th Aerospace Sciences Meeting,
AIAA Paper 78-219, Jan. 1978.
[10] Beckstead, M. W., An Overview of Combustion Mechanisms and
Flame Structures for Advanced Solid Propellants, 36th AIAA/ASME/
SAE/ASEE Joint Propulsion Conference, Huntsville, Al, AIAA
Paper 2000-3325, July 2000.
[11] Chakarvarty, S. R., Price, E. W., Sigman, R. K., and Seitzman, J. M.,
Plateau Burning Behavior of Ammonium Perchlorate Sandwiches and
Propellants at Elevated Pressures, Journal of Propulsion and Power,
Vol. 19, No. 1, Jan.Feb. 2003, pp. 5665.
[12] Fredrick, R. A., Wide Distribution Propellants, U.S. Air Force Rocket
Propulsion Lab., Rept. AFAL-TR-88-073, Edwards AFB, CA,
June 1988.
[13] Iqbal, M. M., and Liang, W., Combustion Calculations of Multimodal
Ammonium Perchlorate Composite Solid Propellants, Theory and
Practice of Energetic Materials, Vol. 6, Science Press, Beijing, China,
2005, pp. 10601068.
[14] Iqbal, M. M., and Liang, W., Propellant Burning Rate Calculations
with Improved Predictions, Journal of Solid Rocket Technology,
Vol. 25, No. 1, 2002.
[15] Beckstead, M. W., and McCarty, K. P., Modeling Calculations for
HMX Composite Propellant, AIAA Journal, Vol. 20, No. 1, 1982,
pp. 106115.
[16] King, M. K., A Model of the Effects of Pressure and Cross Flow
Velocity on Composite Propellant Burning Rate, AIAA 15th AIAA/
ASME/SAE/ASEE Joint Propulsion Conference, Las Vegas, NV,
AIAA Paper 79-1171, June 1979.
[17] Miller, R. R., Control of Solid Distribution in HTPB Propellants, U.S.
Air Force Rocket Propulsion Lab., Rept. TR-78-14, Edwards AFB,
CA, 1978.
[18] Miller, R. R., Effects of Particle Size on Reduced Smoke Propellant
Ballistics, 18th AIAA/ASME/SAE/ASEE Joint Propulsion Conference, Cleveland, OH, AIAA Paper 82-1097, 1982.
[19] Iqbal, M. M., and Liang, W., Combustion Modeling and Formulation
Design of Composite Solid PropellantsExperimental Study, 5th
International Symposium on Multiphase Flow, Heat Mass Transfer and
Energy Conversion [CD-ROM], Xian Jiaotong Univ. Press, Xian,
China, July 2005.

S. Son
Associate Editor

Das könnte Ihnen auch gefallen